cond-mat0504018/p.tex
1: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2: %% Fichier Latex de Michel Pleimling                                          %%
3: %%                                                                            %%
4: %% Version 0.9 -0                                                 12.03.2005  %%
5: %%                                                                            %%
6: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
7: \documentclass[printer]{gPHT2e}
8: \newcommand{\BEQ}{\begin{equation}}     % Gleichungen Anfang ..
9: \newcommand{\BEA}{\begin{eqnarray}}
10: \newcommand{\EEQ}{\end{equation}}       % .. und Ende
11: \newcommand{\EEA}{\end{eqnarray}}
12: \def\be{\begin{equation}}
13: \def\ee{\end{equation}}
14: \def\bc{\begin{center}}
15: \def\ec{\end{center}}
16: \newcommand{\D}{{\rm d}}
17: 
18: \input epsf.sty
19: 
20: 
21: 
22: \begin{document}
23: 
24: 
25: \doi{10.1080/014115940xxxxxxxxxxxx}
26:  \issn{1029-0338}
27: \issnp{0141-1594} \jvol{00} \jnum{00} \jyear{2005}
28: \jmonth{January}
29: 
30: 
31: 
32: 
33: \title{Microcanonical analysis of small systems}
34: \author{Michel Pleimling$^{1}$ and Hans Behringer$^{1,2}$}
35: 
36: 
37: 
38: 
39: 
40: \received{\today}
41: 
42: 
43: \maketitle
44: 
45: 
46: \affil{$^1$ Institut f\"ur Theoretische Physik I,
47: Universit\"at Erlangen-N\"urnberg, D -- 91058 Erlangen, Germany}
48: 
49: \affil{$^2$ Fakult\"at f\"ur Physik, Universit\"at Bielefeld, D -- 33615
50: Bielefeld, Germany}
51: 
52: 
53: 
54: 
55: 
56: 
57: 
58: \begin{abstract}
59: The basic quantity for the description of the statistical
60: properties of physical systems is the density of states or
61: equivalently the microcanonical entropy. Macroscopic quantities of
62: a system in equilibrium can be computed directly from the
63: entropy. Response functions such as the susceptibility
64: are for example related to the curvature of the
65: entropy surface. Interestingly,
66: physical quantities in the microcanonical ensemble show
67: characteristic properties of phase transitions already in 
68: finite systems. In this paper we investigate these characteristics 
69: for finite Ising systems. The
70: singularities in microcanonical quantities which announce a
71: continuous phase transition in the infinite system are characterised by
72: classical critical exponents. Estimates of the non-classical
73: exponents which emerge only in the thermodynamic limit can nevertheless be
74: obtained by analyzing effective exponents or by applying a microcanonical
75: finite-size scaling theory. This is explicitly demonstrated for two-
76: and three-dimensional Ising systems. 
77: 
78: 
79: \end{abstract}
80: 
81: \section{Introduction}
82: In the study of the static properties of finite systems the main quantity
83: of interest is the density of states $\Omega$ as it contains the complete 
84: physical information of the investigated system \cite{Gro01b}. To be specific, consider
85: a magnetic system where $\Omega = \Omega(E,M)$ is a function of the energy $E$
86: and of the magnetization $M$ (which can be a scalar or a vector quantity, depending
87: on the system). With the density of states at hand one may proceed
88: in the canonical way by computing the free energy from which other 
89: quantities such as the susceptibility of the specific heat can be derived.
90: In the last years, however, an increasing number of groups have chosen
91: to directly investigate the density of states or, equivalently, the
92: microcanonical entropy $S =  \ln \Omega$ (setting $k_B=1$).
93: The origin of this increasing interest in the microcanonical approach
94: can be found in the observation that ensembles are not always equivalent
95: \cite{Gro01b,thirr70,bixon89,Hue94a,elli00,daux00,Bar01,Isp01,gulm02}.
96: The inequivalenve of ensemble is for example encountered in situations
97: where the macroscopic limit (for example due to the presence of long-range
98: interactions such as the gravitational force) does not exist.
99: But also in finite systems with short-range interactions the canonical
100: and the microcanonical ensembles may give different results although
101: the ensembles have to become equivalent in the infinite volume limit. This is
102: especially the case for finite systems which in the macroscopic limit
103: undergo a phase transition. It is the purpose of this paper to discuss 
104: the finite-size signatures of phase transitions in the microcanonical
105: ensemble \cite{kast00,Gro00b,huel02,plei04,hove04,behr05,rich05}.
106: 
107: 
108: The paper is organized in the following way. In Section 2 we recall
109: that in the microcanonical approach all quantities of interest
110: can exclusively be expressed as derivatives of the entropy \cite{Gro01b}. In
111: Section 3 we briefly discuss the case of discontinuous phase
112: transitions which lead to a characteristic behaviour of the
113: specific heat in finite microcanonical systems. Section 4 is the 
114: main part of this paper. It is devoted to the signatures of
115: continuous phase transitions encountered in the microcanonical
116: analysis of small systems. We also discuss different ways for
117: extracting the values of critical exponents 
118: directly from the derivatives of the microcanonical entropy. Conclusions
119: are given in Section 5.
120: 
121: 
122: 
123: 
124: \section{Microcanonically defined quantities}
125: 
126: We use in the following the language of magnetism and consider a magnetic system
127: with $N$ degrees of freedom isolated from any environment. The state of the
128: system is then characterized by its energy $E$ and its magnetization $M$.
129: 
130: 
131: 
132: In the microcanonical analysis one considers the microcanonical entropy 
133: density
134: \begin{equation} \label{eq:s}
135: s(e,m) = \frac{1}{N} S(eN,mM) = \frac{1}{N} \ln \Omega (eN,mM) 
136: \end{equation}
137: as function of the energy density $e = E/N$ and the magnetization density
138: $m=M/N$. The entropy surface $s(e,m)$ still exhibits a weak size
139: dependence which is not explicitly denoted here. Sometimes one is only interested in the energy dependence
140: and therefore studies the reduced specific entropy
141: \begin{equation} \label{eq:sR}
142: s_R(e)= \frac{1}{N} \ln \Omega_R (eN)
143: \end{equation}
144: with 
145: \begin{equation}\label{eq:omegaR}
146: \Omega_R (E) = \sum\limits_M \Omega (E,M).
147: \end{equation}
148: Microcanonical quantities can be expressed by  functions of derivatives
149: of the specific microcanonical entropy. For example, the microcanonical
150: temperature reads
151: \begin{equation} \label{eq:T}
152: T = \left( \frac{\D s_R(e)}{\D e} \right)^{-1},
153: \end{equation}
154: which leads to the following expression for the specific heat:
155: \begin{equation} \label{eq:c}
156: c(e)= - \frac{\left( \frac{\D s_R(e)}{\D e} \right)^2}{\frac{\D^2s_R(e)}{\D e^2}}.
157: \end{equation}
158: It is worth noting that the microcanonical temperature and specific heat
159: can also be defined through the entropy density (\ref{eq:s}). These differently
160: defined quantities get identical in the infinite volume limit. We refer the
161: reader to \cite{behr05} for a thorough discussion of this point.
162: 
163: Further quantities discussed in the following include the spontaneous
164: magnetization and the susceptibility. The spontaneous magnetization $m_{sp,N}(e)$ is
165: defined to be the value of $m$ where the entropy (\ref{eq:s}) at a fixed value of $e$
166: has its maximum with respect to $m$ \cite{kast00,huel02,behr05,behr04}:
167: \begin{equation} \label{eq:msp}
168: m_{sp,N}(e) :\Longleftrightarrow s(e, m_{sp,N}(e)) = \max_m s(e,m).
169: \end{equation}
170: To better understand this definition consider the density
171: of states $\Omega(eN,mN)= \exp (N s(e,m))$ for a given energy $e$. The density of
172: states exhibits a sharp maximum as the overwhelming majority of accessible states belongs
173: to the value $m_{sp,N}$ of $m$ where $s$ has its maximum. This is the reason 
174: why we identify $m_{sp,N}$ with
175: the spontaneous magnetization of the finite system.
176: This definition of the microcanonical spontaneous magnetization ensures that we recover the
177: canonically defined spontaneous magnetization in the macroscopic limit.
178: Note also that one can define a magnetic field for the
179: microcanonical ensemble which appears as the conjugate variable to
180: $m$. At the spontaneous magnetization $m_{sp,N}$ the
181: associated magnetic field becomes zero, see \cite{kast00} for further details.   
182: 
183: 
184: Of further interest is the suceptibility \cite{kast00}
185: \begin{equation} \label{eq:chi}
186: \chi(e,m)=- \frac{\frac{\partial s}{\partial e} \frac{\partial^2 s}{\partial e^2}}{\frac{
187: \partial^2 s}{\partial e^2} \frac{\partial^2 s}{\partial m^2} - \left(
188: \frac{\partial^2 s}{\partial e \partial m} \right)^2 }
189: \end{equation}
190: where the denominator on the right hand side is in fact the local curvature of
191: the entropy surface. As it stands, Eq. (\ref{eq:chi}) is only valid in cases where
192: the order parameter is a scalar. For an order parameter with vector character
193: (as encountered for example in classical spin models like the $XY$ or the
194: Heisenberg model) the susceptibility is given by a matrix
195: \cite{behr04b,rich05}. The entries of the susceptibility matrix are
196: then dependent on the local curvature with respect to the different order
197: parameter components.
198: 
199: Closing this Section, we reemphasize that the microcanically defined
200: quantities all fulfill the important request that they become identical to
201: the better known canonical quantities in the infinite system after the
202: trivial transformation from the natural variables $e$ and $m$ of the
203: microcanonical description to the natural variables temperature $T$
204: and magnetic field $h$ of the canonical approach.
205: 
206: \section{Discontinuous phase transitions}
207: The interest in the microcanonical analysis really started with the observation
208: that discontinuous phase transitions display typical and very
209: interesting finite-size signatures in the microcanonical treatment \cite{bixon89,Hue94a,Gro01b}. 
210: Indeed, a discontinuous phase transition
211: reveals itself in finite systems by a convex intruder in the microcanonical
212: entropy which originates from states of coexistence of different phases.
213: In the canonical treatment these states are unstable and cannot be accessed by
214: equilibrium methods. Interestingly, this convex intruder leads to
215: a characteristic backbending of the caloric curve and thus to a negative specific
216: heat. The appearance of a negative specific heat may look like an oddity of
217: the microcanonical approach. However, there have been recent claims that 
218: negative specific heats had been measured (even so in a somewhat indirect way)
219: in experiments on nuclear fragmentation \cite{Dag00} or on the solid-liquid transition
220: of atomic clusters \cite{Sch01a}.
221: 
222: \section{Continuous phase transitions}
223: The main intention of this paper is to discuss the typical finite-size
224: signatures of continuous phase transitions encountered in the microcanonical
225: approach. We thereby focus on classical spin systems and shall discuss in the
226: following the two- and three-dimensional Ising models on hypercubic
227: lattices with periodic boundary conditions. The Ising model is defined by the Hamiltonian
228: \begin{equation} \label{eq:ising}
229: {\mathcal H} = - \sum\limits_{\langle i,j \rangle} \, S_i S_j
230: \end{equation}
231: where the lattice site $i$ is characterized by the classical variable $S_i = \pm 1$.
232: The sum in Equation (\ref{eq:ising})
233: is over nearest neighbour pairs. The two-dimensional three-state
234: Potts \cite{behr04,hove04,behr05} and the three-dimensional $XY$ \cite{rich05} models have also been studied 
235: recently in the present context, but we will
236: not discuss these models here. All the mentioned models have in common that the
237: corresponding infinite system displays a continuous phase transition separating
238: a paramagnetic phase at high temperatures (energies) from a ferromagnetically
239: ordered phase at low temperatures (energies).
240: \subsection{Signatures in small systems}
241: In the canonical ensemble the typical features of spontaneous symmetry breaking
242: like the abrupt onset of the order parameter or a diverging susceptibility
243: are only encountered in the infinite system. In any finite system one does not
244: observe diverging quantities but only rounded maxima. This is, however,
245: completely different in the microcanonical ensemble where the features
246: of spontaneous symmetry breaking turn up already for finite systems
247: \cite{kast00,huel02,behr04,plei04,behr05,rich05,behr03}.
248: 
249: It is instructive to first inspect directly the entropy surface $s(e,m)$.
250: Figure \ref{fig_1} displays this entropy surface for a three-dimensional 
251: Ising model with $6^3$ spins. The fully magnetized states at $e=-3$ and $m= \pm 1$
252: form the two degenerate gound states, whereas the fully disordered macrostate
253: at $e=m=0$ has the highest degeneracy of all. This entropy surface has been
254: computed by a very efficient numerical method based on the concept of
255: transition variables \cite{huel02}. More insights are gained by looking at cuts through
256: the entropy surface along the magnetization at a fixed energy $e$, see Figure \ref{fig_2}. 
257: One observes that the entropy exhibits
258: only one maximum at $m=0$ for energies $e$ above an energy $ e_{c,N}$, whereas for energies $e < e_{c,N}$ two
259: maxima are seen at nonzero magnetizations, separated by
260: a minimum at $m=0$. At the energy $e_{c,N}$ the curvature in
261: magnetization direction vanishes at $m=0$.
262: 
263: \begin{figure}
264: \centerline{\epsfxsize=6.0in\ \epsfbox{figure1.eps}} 
265: \caption{Entropy surface $s(e,m)$ of the three-dimensional Ising model with
266: $N=6^3$ spins as function of the energy $e$ and of the
267: magnetization $m$.} \label{fig_1}
268: \end{figure}
269: 
270: \begin{figure}
271: \centerline{\epsfxsize=4.0in\ \epsfbox{figure2.eps}} 
272: \caption{Cut through the entropy surface at various fixed values of $e$. For $e \geq e_{c,N}$
273: the entropy exhibits a maximum at $m=0$. For $e < e_{c,N}$, two maxima
274: at nonzero magnetizations are visible. At $e = e_{c,N}$ (dashed line) the extremum
275: at $m=0$ changes from a maximum at higher energies to a minimum at
276: lower energies so that the curvature changes its sign.} \label{fig_2}
277: \end{figure}
278: 
279: 
280: Recalling that the spontaneous magnetization is identified with
281: the magnetization where for a fixed energy the density of states is the largest,
282: one can directly extract the value of the spontaneous magnetization at a 
283: given energy from cuts like those shown in Figure \ref{fig_2}.
284: The resulting behaviour of the spontaneous magnetization
285: is discussed in Figure \ref{fig_3} 
286: for three-dimensional Ising models containing $6^3$ and
287: $8^3$ spins. Two different regimes are observed: for $e \geq e_{c,N}$
288: the spontaneous magnetization is zero, whereas for $e < e_{c,N}$
289: it is given by $m = \pm m_{sp,N}$, in accordance with the appearance
290: of two maxima of the entropy surface at lower energies.
291: At the well-defined size-dependent energy $e_{c,N}$ the order
292: parameter sets in abruptly already in finite systems. The energy
293: $e_{c,N}$ may therefore be called transition energy.
294: 
295: \begin{figure}
296: \centerline{\epsfxsize=4.0in\ \epsfbox{figure3.eps}} 
297: \caption{Spontaneous magnetization $m_{sp,N}$ as function of the energy $e$ 
298: measured in three-dimensional Ising models with $6^3$ and $8^3$ spins.
299: At the size-dependent energy $e_{c,N}$ a sharp onset of the order parameter is
300: observed.} \label{fig_3}
301: \end{figure}
302: 
303: Another remarkable finite-size signature of continuous phase transitions is
304: shown in Figure \ref{fig_4}. A divergence of the magnetic susceptibility is
305: observed for finite systems at the same energy $e_{c,N}$ at which the
306: order parameter bifurcates. This divergence is in fact readily understood
307: by recalling, see Equation (\ref{eq:chi}), that the susceptibility is proportional
308: to the inverse of the curvature of the entropy surface. It is the existence
309: of a point of vanishing curvature that is responsible for the observed
310: divergence of the susceptibility, see Figure \ref{fig_2}.
311: 
312: \begin{figure}
313: \centerline{\epsfxsize=4.0in\ \epsfbox{figure4.eps}} 
314: \caption{Magnetic susceptibility $\chi(e,m_{sp,N})$ as function of the energy $e$ 
315: measured in three-dimensional Ising models with $6^3$ and $8^3$ spins.
316: The microcanonically defined reponse function diverges at $e_{c,N}$ already
317: in finite systems.} \label{fig_4}
318: \end{figure}
319: 
320: The microcanonical specific heat, on the other hand, does not diverge in finite 
321: systems. 
322: It nevertheless also displays an intriguing finite-size behaviour,
323: as revealed recently for models with a diverging specific heat in 
324: the  infinite system \cite{behr05}. Indeed, in the microcanonical ensemble
325: the maximum of the specific heat of the corresponding finite systems varies non-monotonically  
326: when increasing the size of the system. Starting with small systems,
327: it first {\it decreases} for increasing system sizes, before it increases
328: for larger systems and finally diverges in the macroscopic limit.
329: This behavious is in strong contrast to the treatment in the canonical ensemble
330: where a monotonic increase of the maximum of the specific heat with the size of the system
331: is always observed \cite{bind72,land76}. In \cite{behr05} 
332: a phenomenological theory has been proposed that acccounts
333: for the observed non-monotonic variation of the maximum of the microcanonical specific heat.
334: 
335: \subsection{Determination of critical exponents directly from the 
336: density of states}
337: At this stage one may wonder whether the typical features of spontaneous symmetry
338: breaking observed in finite microcanonical systems are described by the true critical
339: exponents of the infinite system. A rapid inspection of the numerical data reveals that
340: this is not the case. Indeed, the onset of the order parameter or the divergence
341: of the suceptibility in finite microcanonical systems are in all cases
342: governed by the classical mean-field exponents \cite{kast00,huel02,behr04}.
343: This is readily understood by noting that the entropy
344: surface can be expanded in a Taylor series in the vicinity of
345: $e_{c,N}$ (for a discrete spin model a suitable continuous function
346: that represents the data has to be chosen). As a 
347: consequence one ends up with a Landau-like theory \cite{behr04} which for the microcanonical
348: entropy of a finite system is in principle exact, thus yielding classical values
349: for the critical exponents. In recent years we have developed different approaches
350: that enable us nevertheless to obtain the true  critical exponents by exclusively
351: analyzing the density of states of finite systems \cite{huel02,behr05}.
352: 
353: Before discussing these approaches for the spontaneous 
354: magnetization in some detail we have first to pause briefly
355: in order to recall that in the entropy formalism considered here critical exponents
356: often differ from their thermal counterparts. This is the case when the specific
357: heat diverges algebraically at the critical point of the infinite system, implying that the critical exponent $\alpha$,
358: governing the power-law behaviour of the canonical specific heat in the vicinity of the
359: critical point, is positive. It is then easily shown that the critical exponents
360: $x_\varepsilon$ appearing in the microcanonical analysis are related to the
361: thermal critical exponents $x$ by $x_\varepsilon=x/(1-\alpha)$ \cite{kast00}. However, when the
362: specific heat does not diverge but presents a cusp-like singularity as $\alpha < 0$,
363: we have $x=x_\varepsilon$ \cite{behr04b,rich05}. In the case of a
364: logarithmically diverging specific heat the thermal and microcanonical
365: exponents are also identical.
366: 
367: Let us now come back to the determination of the critical exponents. As already
368: mentioned we obtain classical values for the critical exponents when expanding
369: the finite-size quantities around the energy $e_{c,N}$. The spontaneous
370: magnetization, for example, displays a root singularity for energies
371: $e < e_{c,N}$ close to $e_{c,N}$:
372: \begin{equation} \label{eq:mexp}
373: m_{sp,N}(e)= A_N \left( e_{c,N} - e \right)^{1/2}
374: \end{equation}
375: where the amplitude $A_N$ depends on the system size. In the infinite system
376: one expands around the true critical point $e_c$, leading to the asymptotic
377: behaviour
378: \begin{equation} \label{eq:mexpinf}
379: m_{sp}(\varepsilon) = A \, \varepsilon^{\beta_\varepsilon}
380: \end{equation}
381: for small $\varepsilon $
382: with $\varepsilon = \frac{e_c - e}{e_c - e_g}$, where $e_g$ denotes the ground state
383: energy per degree of freedom. The critical exponent $\beta_\varepsilon$ is given by
384: $\beta_\varepsilon = \beta/(1-\alpha)$ with $\beta$ being the usual thermal critical
385: exponent characterizing the singularity of the spontaneous magnetization. Here we suppose that $\alpha \geq 0$, as it is the case for the Ising
386: model. 
387: 
388: Away from the critical point the spontaneous magnetization of the infinite system
389: is not any more given by the simple expression (\ref{eq:mexpinf}), as correction
390: terms get important. One way to deal with these correction terms is to analyse 
391: the energy dependent effective exponents
392: \begin{equation} \label{eq:beff}
393: \beta_{\varepsilon,eff}(\varepsilon) = \frac{\D \ln m_{sp}(\varepsilon)}{\D \ln \varepsilon}
394: \end{equation}
395: that yields the true critical exponent $\beta_\varepsilon$ in the limit $\varepsilon 
396: \longrightarrow 0$ \cite{huel02}.
397: 
398: In order to use Equation (\ref{eq:beff})
399: we need data free of finite-size effects. On the other hand one observes in
400: Figure \ref{fig_3} that there exists a energy interval for which the spontaneous
401: magnetizations computed in systems of different sizes are identical.
402: This suggest the following procedure for computing $\beta_\varepsilon$.
403: Calculate the spontaneous magnetization for a series of increasing system sizes $\left\{ N_j , j=1, \cdots ,  n \right\}$.
404: In the energy range where $N_i$ and $N_{i+1}$ yield the same exponent 
405: $\beta_{\varepsilon,eff}$, select this exponent for the infinite system, until finite-size
406: effects show up. Then plot the common exponent for system sizes $N_{i+1}$ and
407: $N_{i+2}$ until these begin to disagree and so on. This procedure is rather cumbersome
408: but ensures that data essentially free of finite-size effects are obtained.
409: 
410: The result of this approach \cite{huel02} for the two- and the three-dimensional Ising models
411: is shown in Figure \ref{fig_4}. For the three-dimensional case we observe that 
412: $\beta_{\varepsilon,eff}$ rapidly approaches a constant value that perfectly agrees with the
413: expected asymptotic value $\beta_\varepsilon = 0.367$. In two dimensions the agreement is not 
414: as perfect, even so a reasonable estimate of $\beta_\varepsilon$ 
415: is obtained by linearly extrapolating the existing data.
416: The situation is indeed more complicated in the two-dimensional Ising model, as one
417: is dealing here with the rather pathological case of a logarithmically diverging specific
418: heat with $\alpha = 0$. In fact, we expect for any model with $\alpha \neq 0$
419: a similar good agreement between the expected value of
420: $\beta_\varepsilon$ and the value obtained from linearly extrapolating $\beta_{eff,\varepsilon}$ as for
421: the three-dimenisonal Ising model. This expectation
422: has recently been confirmed for the three-dimensional $XY$ model \cite{rich05}.
423: 
424: \begin{figure}
425: \centerline{\epsfxsize=4.0in\ \epsfbox{figure5.eps}} 
426: \caption{Effective exponent $\beta_{\varepsilon,eff}$ versus reduced
427: energy $\varepsilon$ for the two- and the three-dimensional Ising models.
428: The dashed lines indicate the expected values of $\beta_\varepsilon$, see text.
429: Error bars are only shown when they are larger than the sizes of the symbols.} \label{fig_5}
430: \end{figure}
431: 
432: The determination of critical exponents via effective exponents has the major
433: drawback that data free of finite-size effects have to be obtained, which means
434: that systems of increasing sizes have to be simulated on approach to the critical 
435: point. For Figure \ref{fig_5} our largest systems contained $700 \times 700$ spins
436: in two and $80 \times 80 \times 80$ spins in three dimensions. Simulating these
437: systems is very time consuming even with the best algorithms available.
438: The second approach for determining the values of critical exponents discussed
439: in the following does not need large systems, but instead it explicitly takes
440: advantage of the finite-size effects.
441: 
442: In the canonical ensemble finite-size scaling theory is a valuable and often
443: used tool for extracting critical exponents from finite-size data. Its starting
444: point is the observation that in the asymptotic limit $L \longrightarrow \infty$ 
445: and $T \longrightarrow T_c$ ($T_c$ being the critical temperature)
446: the behaviour of finite-size quantities is governed
447: by scaling functions that are basically determined by the ratio
448: $L/\xi(T)$ with $\xi(T)$ being the correlation length and $L=N^{1/d}$
449: the linear extension of the $d$-dimensional system \cite{Bar83}.
450: %A similar approach was lacking
451: %for the microcanonical ensemble until recently. 
452: In \cite{plei04} we have developed
453: a microcanonical finite-size scaling theory that takes advantage of
454: the existence of a well-defined transition point $e_{c,N}$ in finite
455: microcanonical systems.
456: 
457: The starting point is the scaling behaviour of the entropy of finite systems
458: considered as a function of the energy, the magnetization and the inverse
459: system size. This then leads to the scaling form
460: \begin{equation}  \label{eq:scal}
461: L^{\beta_\varepsilon/\nu_\varepsilon} m_{sp,N}\left( e_{c,N} - e \right) 
462: \approx W\left( C ( e_{c,N} - e) L^{1/\nu_\varepsilon} \right)
463: \end{equation}
464: of the spontaneous magnetization (\ref{eq:msp}). Here $W$ is a scaling
465: function characterizing the given universality class and $C$ is a nonuniversal
466: constant which is different for the various model systems belonging to
467: the same universality class. The exponent $\nu_\varepsilon$ is given by $\nu_\varepsilon= \nu/(1-\alpha)$
468: where $\nu$ is the usual thermal critical exponent governing the divergence
469: of the correlation length on approach to the critical point.
470: As the spontaneous magnetization presents a square root singularity
471: in any finite microcanonical system, the scaling function $W$ varies for
472: small scaling variables $x=C ( e_{c,N} - e) L^{1/\nu_\varepsilon}$ as
473: $W(x) \sim \sqrt{x}$. In Figure \ref{fig_6} we test the scaling form (\ref{eq:scal})
474: for the usual Ising model (\ref{eq:ising}) with only nearest neighbour interactions
475: (denoted by I)
476: and for a generalized Ising model with equivalent nearest and next-nearest
477: neighbour interactions (denoted by GI). The Hamiltonian of this latter model is given by
478: \begin{equation} \label{eq:gi}
479: {\mathcal H}_{GI} = - \sum\limits_{\langle i,j \rangle} \, S_i S_j
480: - \sum\limits_{\left( i,k \right)} \, S_i S_k
481: \end{equation}
482: where the second sum is over bonds connecting next-nearest neighbours.
483: The model (\ref{eq:gi}) belongs to the same universality class as the standard
484: Ising model (\ref{eq:ising}) so that both models should yield the same values
485: for universal quantities.
486: 
487: \begin{figure}
488: \centerline{\epsfxsize=4.0in\ \epsfbox{figure6.eps}} 
489: \caption{Microcanonical finite-size scaling plot for the Ising model with only
490: nearest neighbour interactions (I) and for the Ising model with equivalent nearest and next-nearest
491: neighbour interactions (GI). Critical exponents are determined by the best data collapse.
492: By adjusting the nonuniversal constant $C$ the data of both models fall on a unique curve,
493: thus demonstrating the universality of the scaling function $W$ given in 
494: Eq.\ (\ref{eq:scal}). Error bars are smaller than the sizes of the symbols.} \label{fig_6}
495: \end{figure}
496: 
497: Figure \ref{fig_6} illustrates two different aspects of the scaling behaviour of the
498: microcanonically defined spontaneous magnetization. On the one hand it shows that both
499: for the Ising model and for the GI model a data collapse can be achieved by plotting
500: $L^{\beta_\varepsilon/\nu_\varepsilon} m_{sp,N}$ as a function of 
501: $( e_{c,N} - e) L^{1/\nu_\varepsilon}$. The values of the involved critical exponents
502: are thereby obtained by the best data collapse. This yields the values
503: $\beta_\varepsilon/\nu_\varepsilon = 0.54 \pm 0.03$ ($0.51 \pm 0.03$) and
504: $1/\nu_\varepsilon = 1.43 \pm 0.01$ ($1.43 \pm 0.01$) for the Ising (GI) 
505: model, in excellent agreement
506: with the expected values $\beta_\varepsilon/\nu_\varepsilon = 0.52$ and
507: $1/\nu_\varepsilon = 1.43$. On the other hand Figure \ref{fig_6} also proves
508: that the scaling function $W$ appearing in Eq.\ (\ref{eq:scal}) is the same for both
509: models, in accordance with its expected universality. Indeed, the data of both models
510: fall on a common master curve when adjusting the nonuniversal constant $C$, so that
511: $C_{GI} = 0.219 C_I$. As shown in \cite{plei04} scaling functions obtained for models belonging
512: to different universality classes are indeed different.
513: 
514: The microcanonical finite-size scaling theory, which has also been
515: applied successfully to the
516: Potts \cite{plei04} and to the $XY$  \cite{rich05} models, is very promising. One of the remarkable
517: point is that no {\it a priori} knowledge of the infinite system is needed. Especially,
518: one does not need to know the exact location of the critical point of the infinite system.
519: 
520: \section{Conclusion}
521: In this paper we have discussed the signatures of phase transitions in finite
522: microcanonical systems. Discontinuous phase transitions are announced
523: in the microcanonical ensemble by a backbending of the caloric curve and the
524: appearance of negative specific heat. Continuous phase transitions also lead
525: to intriguing finite-size signatures. Indeed, typical features of spontaneous
526: symmetry breaking like the onset of the order parameter or the divergence of
527: the susceptibility are already encountered  in finite microcanonical systems.
528: These singularities are governed by the classical mean-field exponents. We have
529: discussed in this paper two different approaches that nevertheless enable us
530: to extract the true critical exponents from the density of states of finite systems.
531: One approach involves effective exponents, whereas the other is based on a
532: microcanonical finite-size scaling theory. The latter approach explicitly
533: takes advantage of the existence of a well-defined transition point in any
534: finite microcanonical system.\\~\\
535: 
536: {\bf Acknowledgements:}\\~\\
537: It is a pleasure to thank Alfred H\"{u}ller for many years of fruitful
538: and enjoyable collaboration. 
539: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
540: \begin{thebibliography}{999}
541: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
542: \bibitem{Gro01b} D. H. E. Gross, {\sl Microcanonical Thermodynamics:
543: Phase Transitions in 'Small' Systems}
544: ({\it Lecture Notes in Physic} 66) (World Scientific, Singapore 2001).
545: \bibitem{thirr70} W. Thirring, Z.\ Phys. {\bf 235}, 339 (1970).
546: \bibitem{bixon89} M. Bixon and J. Jortner, J.\,Chem.\,Phys. {\bf 91},
547: 1631 (1989).
548: \bibitem{Hue94a} A. H\"{u}ller, Z.\ Phys. B {\bf 93}, 401 (1994).
549: \bibitem{elli00} R. S. Ellis, K. Haven and B. Turkington,
550: J.\ Stat.\ Phys. {\bf 101}, 999 (2000).
551: \bibitem{daux00} T. Dauxois , P. Holdsworth and S. Ruffo,
552: Eur.\ Phys.\ J. B {\bf 16}, 659 (2000).
553: \bibitem{Bar01} J. Barr\'{e}, D. Mukamel and S. Ruffo,
554: Phys. Rev. Lett. {\bf 87}, 030601 (2001).
555: \bibitem{Isp01} I. Ispolatov and E. G. D. Cohen,
556: Physica A {\bf 295}, 475 (2001).
557: \bibitem{gulm02} F. Gulminelli and Ph. Chomaz, Phys. Rev. E {\bf 66}, 046108 (2002).
558: \bibitem{kast00} M. Kastner, M. Promberger and A. H\"{u}ller, J. Stat. Phys.
559: {\bf 99}, 1251 (2000).
560: \bibitem{Gro00b} D. H. E. Gross and E. V. Votyakov,
561: Eur.\ Phys.\ J. B {\bf 15}, 115 (2000).
562: \bibitem{huel02} A. H\"{u}ller and M. Pleimling, Int. J. Mod. Phys. C {\bf 13}, 947 (2002).
563: \bibitem{plei04} M. Pleimling, H. Behringer and A. H\"{u}ller, Phys. Lett. A
564: {\bf 328}, 432 (2004).
565: \bibitem{hove04} J. Hove, Phys.\ Rev.\ E {\bf 70}, 056707 (2004).
566: \bibitem{behr05} H.\ Behringer, M.\ Pleimling and A.\ H\"{u}ller,
567: J. Phys. A: Math.\ Gen.\ {\bf 38}, 973 (2005).
568: \bibitem{rich05} A. Richter, M.\ Pleimling and A.\ H\"{u}ller, cond-mat/0503733.
569: \bibitem{behr04} H. Behringer, J. Phys. A: Math. Gen. {\bf 37}, 1443 (2004).
570: \bibitem{behr04b} H.\ Behringer, {\it On the structure of the entropy surface
571: of microcanonical systems} (Mensch und Buch Verlag, Berlin, 2004).
572: \bibitem{Dag00} M. D'Agostino, F. Gulminelli, Ph. Chomaz, M. Bruno,
573: F. Cannata, R. Bougault, F. Gramegna, I. Iori, N. Le Neindre, G. V. Margagliotti,
574: A. Moroni and G. Vannini, Phys.\ Lett. B {\bf 473}, 219 (2000).
575: \bibitem{Sch01a} M. Schmidt, R. Kusche, T. Hippler, J. Donges, W. Kronm\"uller, B. von Issendorff,
576: and H. Haberland, Phys. Rev. Lett. {\bf 86}, 1191 (2001).
577: \bibitem{behr03} H. Behringer, J. Phys. A: Math. Gen. {\bf 36}, 8739 (2003).
578: \bibitem{bind72} K. Binder, Physica {\bf 62}, 508 (1972).
579: \bibitem{land76} D. P. Landau, Phys. Rev. B {\bf 13}, 2997 (1976).
580: \bibitem{Bar83} M.\,N.\,Barber, in {\it Phase Transitions and Critical
581: Phenomena}, vol~8,
582: edited by C.~Domb and J.~L.~Lebowitz  (Academic Press, London and New York, 1983).
583: 
584: \end{thebibliography}
585: 
586: \end{document}
587: